Kinetics and Structural Changes in CO2 Capture of K2CO3 under a

Jun 17, 2015 - The capacity and kinetics of CO2 capture of K2CO3 were studied to determine the mechanism for CO2 sequestration under ambient condition...
0 downloads 0 Views 5MB Size
Subscriber access provided by NEW YORK UNIV

Article

Kinetics and Structural Changes in CO2 Capture of K2CO3 under a Moist Condition Hongchao Luo, Hideyuki Chioyama, Stephan Thürmer, Tomonori Ohba, and Hirofumi Kanoh Energy Fuels, Just Accepted Manuscript • DOI: 10.1021/acs.energyfuels.5b00578 • Publication Date (Web): 17 Jun 2015 Downloaded from http://pubs.acs.org on June 21, 2015

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Energy & Fuels is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

Kinetics and Structural Changes in CO2 Capture of K2CO3 under a Moist Condition Hongchao Luo,a Hideyuki Chioyama,a Stephan Thürmer,b Tomonori Ohba,a and Hirofumi Kanoha* a

Graduate School of Science, Chiba University, 1–33 Yayoi-Cho, Inage-Ku, Chiba, Japan

b

Center for Frontier Science, Chiba University, 1–33 Yayoi-Cho, Inage-Ku, Chiba, Japan

ABSTRACT: The capacity and kinetics of CO2 capture of K2CO3 were studied to determine the mechanism for CO2 sequestration under ambient conditions. Bicarbonate formation of K2CO3 was examined by a thermogravimetric analysis under various CO2 concentrations in the presence of water vapor and the accompanying structural changes of K2CO3 were demonstrated by X-ray diffraction (XRD). Morphological variations were observed during the reaction in the presence of different CO2 concentrations through scanning electron microscope (SEM). Structural changes and morphological variations, which occurred during the course of the reaction, were then connected to the kinetic and exothermic properties of the CO2 capture process from XRD and SEM measurements. The XRD results showed that the bicarbonate formation process of K2CO3 could be divided into three reactions such as the formation of K2CO3·1.5H2O from K2CO3, the subsequent formation of K4H2(CO3)3·1.5H2O from K2CO3·1.5H2O, and the slow formation of KHCO3 from K4H2(CO3)3·1.5H2O. The SEM observations showed that the morphology of the

ACS Paragon Plus Environment

1

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 24

particles at all three stages played a crucial role in the kinetic behavior for CO2 sorptivity of K2CO3. CO2 capture of K2CO3 was inhibited under a concentrated CO2 atmosphere during the initial stage consisting of the first and second reactions, but the formation of KHCO3 from K4H2(CO3)3·1.5H2O was thermodynamically favorable upon the increase of the CO2 concentration.

1. INTRODUCTION As the rapid increase in CO2 emissions from human activities is considered to be the main cause of global warming, reduction of CO2 emissions has become an urgent global issue. Carbon capture and storage (CCS) can provide an immediate solution by stabilizing or reducing the atmospheric CO2 concentration via development of CO2 sequestration methods and accompanying renewable energy technologies.1–5 The main focus of this field has been placed on amine-, Zeolite13X-, and activated carbon-based, CO2 capture.6–8 However, these technologies still face challenges, such as high energy consumption during sorbent regeneration, amine loss during regeneration, and low amine utilization efficiency. Furthermore, Zeolite13X- and activated carbon-based CO2 sorbents under ambient conditions have CO2 efficiency issues because of interference between the adsorbed CO2 and H2O, as well as high cost for gas removal. Furthermore, these sorbents are not appropriate for CO2 capture applications under moist conditions. Feasible CO2 capture sorbents must provide significant enhancement in the capture or sorption rate, saturated sorption amount, and stability for long-term CO2 capture. Zhao et al. discovered that K2CO3 derived from KHCO3 showed excellent CO2 capture capacity. This can be used to solve the problem of an overall low CO2 sorption rate.9 The practical application of K2CO3 as a

ACS Paragon Plus Environment

2

Page 3 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

CO2 sorbent has already been demonstrated on a pilot scale in an actual coal-fired power plant.10,11 Conversely, K2CO3 as a CO2 sorbent under moist conditions at atmospheric pressure has been studied extensively as the following reaction:9, 12–21 K2CO3(s) + CO2(g) + H2O(g) ⇆ 2 KHCO3(s)

(1)

where the forward, exothermic reaction (termed “bicarbonate formation”) is the formation of KHCO3 (the theoretical CO2 capture amount of K2CO3: 7.24 mmol·g–1), while the reverse, endothermic reaction is the decomposition of KHCO3. Lower temperature increases the CO2 sorption amount of K2CO3 according to the forward reaction of (1); thus, K2CO3 may be used at room temperature for a highly saturated sorption value. Moreover, Chioyama et al. found that the occlusion amount (6.48 mmol·g–1, 285 mg·g–1) at 0.1 MPa and 313 K was superior to that at higher temperature and showed that K2CO3 could be used at room temperature.22 Furthermore, the favorable CO2 occlusion performance of K2CO3 supported on activated carbon at low temperature has also been demonstrated by enhancing the surface area.23 Consequently, K2CO3 is especially promising as a CO2 sorbent material for moist conditions at room temperature. However, most previous studies regarding CO2 capture of K2CO3 were carried out at relatively high temperatures (>333 K) under atmospheric pressure. In numerous studies on the CO2 sorption of K2CO3 under moist conditions, CO2 capture did not always proceed directly via reaction (1). Hayashi et al. reported significantly improved CO2 capture capacity using K2CO3·1.5H2O as the active species under flue gas conditions of 13.8% CO2 with 10% H2O in helium at 373 K.18 Shigemoto et al. proposed that K2CO3 supported on activated carbon could be converted to KHCO3 through the reaction of K2CO3·1.5H2O with CO2 under moist flue gas (10 % H2O, 11.8 % CO2) at 363 K. The relevant reactions are:19

ACS Paragon Plus Environment

3

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 24

K2CO3(s) + 1.5H2O(g) ⇄ K2CO3·1.5H2O(s)

(2)

K2CO3·1.5H2O(s) + CO2(g) ⇄ KHCO3(s) + 0.5H2O(g)

(3)

Bicarbonate formation of K2CO3 on Al2O3 was achieved via reactions (2) and (3) rather than reaction (1), as reported by Seo et al., 20 and performed under (10 vol. % CO2, 12.2 vol. % H2O, 77.8 vol. % N2)20 atmosphere within a temperature range of 323–343 K. In addition, Lee et al.21 presented similar bicarbonate formation of K2CO3 on Al2O3 under (1 vol. % CO2, 9 vol. % H2O and 90 vol. % N2)21 atmosphere at above 323 K. However, Zhao et al. suggested that the formation of K4H2(CO3)3·1.5H2O may also play an important role in bicarbonate formation of K2CO324 and would proceed via reactions (4) and (5) instead of reaction (1) in an atmosphere of 15 mol % CO2, 15 mol % H2O and 70 mol% N2 at 333 K:25 2K2CO3(s) + 2.5H2O(g) + CO2(g) ⇄ K4H2(CO3)3·1.5H2O(s)

(4)

K4H2(CO3)3·1.5H2O(s) + CO2(g) ⇄ 4KHCO3(s) + 0.5H2O(g)

(5)

These reactions were shown to proceed differently depending on the experimental conditions. As such, the different kinetics of CO2 sorption reactions of K2CO3 were demonstrated in these works based on a mechanistic interpretation. However, no detailed report on temperature and crystal structure changes with reaction time has been conducted at low temperatures. The morphologies of K2CO3 particles also change as the reactions advance, as the crystal growth depends on the reaction conditions. This information is also important for understanding the kinetic behavior of K2CO3 particles. However, details of the morphological variations with reaction time have not been reported; thus, the CO2 capture mechanism under various CO2 and H2O conditions is currently unclear.

ACS Paragon Plus Environment

4

Page 5 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

In this paper, CO2 sorptivity of K2CO3 under various CO2 concentrations in the presence of saturated water vapor was examined by thermogravimetric-differential thermal analysis (TG– DTA) and accompanying structural changes with reaction time were investigated using X-ray diffraction (XRD) method under vacuum. In addition, morphological variations with reaction time were observed using a scanning electron microscope (SEM). Based on the crystal structure and kinetic analysis of bicarbonate formation of K2CO3, three relevant reactions are proposed as fundamental pathways in the KHCO3 formation from K2CO3 at low temperatures. 2. EXPERIMENTAL SECTION 2.1. Sample preparation. Analytical reagent-grade KHCO3 (99.5 % chemical purity, Wako Chemical Co., Ltd.) was used throughout the decomposition experiments of KHCO3 and bicarbonate formation measurements of K2CO3. The decomposition of KHCO3 was processed with TG–DTA (Shimadzu, DTG–60AH). KHCO3 (44–50 mg) was heated at 5 K·min–1 from 298 K to 473 K in an atmosphere of pure N2 at 100 cm3·min–1 to form K2CO3 and held at these conditions for 5 min to complete the decomposition. Subsequently, the temperature was lowered to 313 K at -5 K·min–1 and maintained for 30 min under a N2 atmosphere. 2.2. Bicarbonate formation measurements. Measurement of the K2CO3 bicarbonate formation with varying reaction duration times was performed with the TG–DTA apparatus with various compositions of CO2 and N2 containing saturated water vapor at 313 K. CO2 and N2 were obtained from high-purity gas cylinders equipped with mass flow controllers to control the flow rate. H2O vapor was produced by flowing N2 and CO2 into distilled water in a bubbler for more than 5 h to reach saturation, with N2 as the balance gas. The mixed CO2, H2O, and N2 gas was then introduced into a moisture detector with a thermometer and hygrometer at 313 K. The

ACS Paragon Plus Environment

5

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 24

relative humidity (RH) of the mixed gas was measured with the hygrometer. The total gas flow rate was maintained at 100 cm3·min–1 for in all experiments. 2.3. Crystal structure and morphology measurements. The crystal structures of the products after reaction times of 1, 5, 20, 40, 60, and 120 min were measured by XRD (MAC Science, M03XHF) under vacuum to avoid occlusion of CO2 or H2O from the ambient atmosphere. The powder XRD patterns were obtained in the 2θ range of 25° to 45°using CuKα radiation (40 kV, 25 mA, λ = 0.15406 nm) at room temperature. The morphology of the particles before and after CO2 capture for the different reaction times was observed by using scanning electron microscope (SEM; JEOL, JSM–6510A) after coating the samples with metallic osmium. 3. RESULTS AND DISCUSSION 3.1. Kinetics in CO2 sorption of K2CO3 3.1.1. Bicarbonate formation of K2CO3. The decomposition of KHCO3 was processed with TG–DTA under a N2 atmosphere via the reverse reaction of (1) to form K2CO3, CO2 and H2O, and the resulting time course of weight is shown in Figure S1. The weight for the decomposition of KHCO3 from 0 min to 101 min gradually decreased to 69 % (± 0.2 %), which is in very good agreement with the theoretical value of 68.9% according to the reverse reaction of (1). This confirms that the decomposition reaction proceeds to completion. In order to examine the effect of CO2 concentration on the sorptivity of K2CO3, TG data for the bicarbonate formation of K2CO3 were obtained under various partial pressures of CO2, as shown for reaction time of 0–120 min in Figure 1. A sufficiently humid atmosphere (RH: 68% at 313 K) was supplied to K2CO3 with a mixed gas of CO2 and N2 containing saturated water vapor.

ACS Paragon Plus Environment

6

Page 7 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

When the flow rate of CO2 was only 1 cm3·min–1 + 99 cm3·min–1 N2 (1%V/V CO2), the sorption amount was 6.24 mmol·g–1, which is 86.2% of the theoretical amount of 7.24 mmol·g–1. As the CO2 flow rate increased from 1 to 30 cm3·min–1, the sorption amount of K2CO3 increased from 6.24 to 6.85 mmol·g–1. According to reaction (1), high CO2 concentration leads to an equilibrium shift toward bicarbonate formation of K2CO3. The sorption amount was 6.92 mmol·g–1, when the CO2 flow rate was 100 cm3·min–1, which is 95.6% of the theoretical value. In comparison with the CO2 sorption amount at a CO2 flow rate of 30 cm3·min–1, the experimental results indicated that increasing the CO2 concentration resulted in a greater conversion of K2CO3 to KHCO3 in the presence of saturated water vapor at 313 K. Figure 2 shows the XRD patterns after the completed occlusion (120 min mark in Figure 1) at various CO2 flow rates. It is evident that the peaks of both the K4H2(CO3)3·1.5H2O and KHCO3 phases were present in all samples, while the peaks of the K4H2(CO3)3·1.5H2O phase appeared to decrease with increasing the CO2 concentration, where were consistent with the TGA data. However, bicarbonate formation was slow at CO2 flow rates of 20–100 cm3·min–1. In contrast, bicarbonate formation at CO2 flow rates of 5 and 10 cm3·min–1 was much faster than other CO2 flow rates, as can be seen in Figure 1b. It was shown that the layer of transformed KHCO3 on the interface of K2CO3 particles inhibited the CO2 and H2O transfer process due to increasing surface resistance increasing.25 On the other hand, the initial sorption rate was limited at CO2 flow rates above 20 cm3·min–1, as can be seen in Figure 1b. It was previously reported that the formation of K2CO3·1.5H2O as an active species could enhance the CO2 capture capacity, with its formation being the limiting step.18, 19 However, Zhao et al. discovered that the initial product after the reaction of K2CO3 with CO2 and H2O was not K2CO3·1.5H2O but rather K4H2(CO3)3·1.5H2O. The bicarbonate formation rate of K2CO3 was much faster than that of K2CO3·1.5H2O.24, 25

ACS Paragon Plus Environment

7

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 24

Hence, the formation of the initial two products, their transformation processes, and the morphological variation on the surface of K2CO3 particles may influence the reaction rate of bicarbonate formation under different CO2 concentrations. Investigation of the transformation process of K2CO3 and the associated changes in temperature during bicarbonate formation is of the utmost importance for a better understanding of the CO2 sorptivity kinetics of K2CO3. 3. 1.2. Exothermic properties and temperature variation. As the transformation process of K2CO3 to bicarbonate leads to a temperature change of the reaction system, the kinetic properties can be tracked by monitoring the temperature during the transformation under various CO2 concentrations. Figure 3 shows the temperature curves (a), DTA curves (b), and derivative thermogravimetric (DTG) curves (c; obtained by differentiating the weight change in Figure 1 over time) for various CO2 concentrations. Both DTA and DTG curves showed a striking similarity, exhibiting a series of peaks at the initial and the final stages (close to equilibrium) corresponding to exothermic reactions of initial product formation from K2CO3 and initial product transformation to KHCO3, respectively. In particular, temperature variation in the initial stage displayed remarkable similarity to both DTA and DTG curves at this stage, as shown in Figure 3d–f. There was a rapid increase in temperature in the initial stage of the reaction (reaction time of < 5 min), as can be seen in Figure 3d. The maximum temperature of this initial stage increased with the CO2 flow rate from 1 to 10 cm3·min–1, but decreased with higher CO2 flow rate. After the initial increase in temperature, the temperature decreased moderately in all curves (middle stage). A second abrupt increase was observed for CO2 flow rates between 20 and 100 cm3·min–1 when the CO2 sorption was near equilibrium, arguably due to a successive exothermic reaction, although no marked peak was observed for CO2 flow rates of 1 and 5 cm3·min–1. This second

ACS Paragon Plus Environment

8

Page 9 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

peak had an analogous peak in the DTA and DTG curves, as shown in Figure 3b and c. The temperature at this stage (final stage) decreased upon increase of the CO2 flow rate from 20 to 100 cm3·min–1. These results indicate that both the overall reaction and the initial process of bicarbonate formation under high CO2 flow rate (above 20 cm3·min–1) proceed slowly, which can be understood together with analysis of the reaction mechanism of the system, as discussed below. 3.2. Structural changes in CO2 sorption of K2CO3. The different pathways of bicarbonate formation of K2CO3 were examined by measuring the XRD patterns of CO2-occluded samples of K2CO3 in the presence of water vapor over various reaction times. There have been few reports related to the initial transformation process of K2CO3 and no detailed mechanisms have been reported so far. Lee et al. reported that K2CO3 was transformed to K4H2(CO3)3·1.5H2O upon treatment with water vapor, followed by further transformation from K4H2(CO3)3·1.5H2O to K2CO3·1.5H2O by drying in N2. 26, 27 It was also claimed that K2CO3·1.5H2O was readily converted to K2CO3 at low temperature under a slightly moist conditions, supported by the heat released from the exothermic reaction.26, 27 However, our results, shown in Figure 4, indicate that the initial product of CO2 occlusion of K2CO3 is not a simple transformation from K2CO3 to K2CO3·1.5H2O and K4H2(CO3)3·1.5H2O. For the reaction occurring over 1 min, the XRD patterns (Figure 4a) showed the presence of two phases, K2CO3·1.5H2O and K4H2(CO3)3·1.5H2O phase, for all CO2 flow rates. K2CO3 was completely converted to these two phases under humid conditions (RH: 68%, 313 ± 3 K). Thus, it can be concluded that the temperature changes are a result of the initial stage and the subsequent transformation processes.

ACS Paragon Plus Environment

9

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 24

Subsequently, when the reaction proceeded beyond 5 min, the peak intensity of K2CO3·1.5H2O and K4H2(CO3)3·1.5H2O in the XRD pattern significantly increased (Figure 4b). Upon increasing the reaction time to 20 min, the dominant peaks were found to originate from the K4H2(CO3)3·1.5H2O phase. However, the peaks of KHCO3 increased only slightly with increasing CO2 concentration. Furthermore, compared with the patterns obtained for the 5 min reaction, the XRD patterns after an occlusion reaction for 20 min showed that the peaks of K4H2(CO3)3·1.5H2O increased considerably in intensity, while those of the K2CO3·1.5H2O phase nearly disappeared. These changes indicate that K2CO3·1.5H2O was converted to K4H2(CO3)3·1.5H2O rather than KHCO3. Accordingly, it can be concluded that K2CO3·1.5H2O is the initial product in the bicarbonate formation process of K2CO3. Next, when the reaction time was extended to 40 min, the peak intensity of the KHCO3 phase increased significantly, while the K4H2(CO3)3·1.5H2O phase was diminished (Figure 4d). This result confirmed that K4H2(CO3)3·1.5H2O was converted to KHCO3 through reaction (5). At a low CO2 flow rate (1 cm3·min–1), it was observed that the K2CO3·1.5H2O phase still remained. This is a result of the unhindered transformation of K2CO3·1.5H2O from K4H2(CO3)3·1.5H2O based on the reversibility of the reaction in a dilute CO2 atmosphere and saturated humidity at low temperature. Conversely, at 100 cm3·min–1, the presence of the K2CO3·1.5H2O phase also demonstrated the slow rate of CO2 occlusion of K2CO3 under such a high CO2 flow. Upon increasing the reaction time to 60 min, Figure 4e shows peaks consisting of K4H2(CO3)3·1.5H2O and KHCO3 phase contributions under a CO2 flow rate of 100 cm3·min–1, while K2CO3·1.5H2O and K4H2(CO3)3·1.5H2O phases were formed under a flow rate of only 1 cm3·min–1. The peaks of the K4H2(CO3)3·1.5H2O phase were found to decrease at the flow rate

ACS Paragon Plus Environment

10

Page 11 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

of 100 cm3·min–1. These results were consistent with those of TGA and the final crystal structure observed under various CO2 concentrations. Based on the discussed transformation processes and kinetics of CO2 sorptivity of K2CO3 under various CO2 concentrations and in the presence of saturated water vapor, we can summarize the bicarbonate formation pathways of K2CO3 as follows instead of reaction (1): K2CO3(s) + 1.5H2O(g) ⇄ K2CO3·1.5H2O(s)

(6)

2K2CO3·1.5H2O(s) + CO2(g) ⇄ K4H2(CO3)3·1.5H2O(s) + 0.5H2 (g)

(7)

K4H2(CO3)3·1.5H2O(s) + CO2(g) ⇄ 4KHCO3 + 0.5H2O(g)

(8)

The crystallite size of K2CO3·1.5H2O, K4H2(CO3)3·1.5H2O, and KHCO3 for each reaction time was estimated from the XRD patterns in Figure 4a–d. The K2CO3·1.5H2O and K4H2(CO3)3·1.5H2O phases were found to grow more slowly with reaction time for high CO2 flow rate (30–100 cm3·min–1) than for low CO2 flow rate (1–5 cm3·min–1), as shown in Figure S2a and b. Furthermore, it was shown visually that the transformation from K4H2(CO3)3·1.5H2O to KHCO3 proceeded more slowly at high CO2 flow rates (30–100 cm3·min–1) than that for low CO2 flow rate (5 cm3·min–1), as shown for 20–40 min in Figure S2c. Thus, it can be confirmed that both the initial process and overall reaction of bicarbonate formation under high CO2 flow rate (above 20 cm3·min–1) proceed more slowly. Both low temperatures and high CO2 concentrations are favorable for an equilibrium shift toward bicarbonate formation, but the initial and final exothermic reaction properties and the CO2 capture rate of K2CO3 under different CO2 concentrations cannot be obtained from XRD data alone, as discussed below.

ACS Paragon Plus Environment

11

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 24

3.3. Sample morphologies with reaction time. To gain an understanding of these properties, the morphological change of K2CO3 was examined via SEM before and after CO2 occlusion for various reaction times. The exemplary images for 5, 30, and 100 cm3·min–1 CO2 atmospheres are shown in Figures 5, 6, and S3. As can be seen in Figure 5a, K2CO3 particles, which were formed by the decomposition of KHCO3 under a N2 atmosphere, had a macroporous surface. For a CO2 flow rate of 5 cm3·min–1 (Figure 5b), which is milder for the bicarbonate formation than the 100 cm3·min–1 flow rate (Figure 3a and 3d), the diameter of the macropores on the surface slightly decreased after the initial 5 min. After 20 min, the pores on the surface nearly disappeared (Figure 5c), corresponding to products consisting of major K4H2(CO3)3·1.5H2O and minor KHCO3 (Figure 4c). Large granules of KHCO3 developed on the external surface (Figure 5d) along with bicarbonate formation of K4H2(CO3)3·1.5H2O reacting with CO2 and H2O for 20– 120 min (Figure 4d and Figure 2). During this transformation period, CO2 occlusion reached a plateau in approximately 40 min (Figure 1), with no second peak developing in the DTG and DTA curves (Figure 3). At a CO2 flow rate of 30 cm3·min–1, as shown in Figure S3, the macropores on the surface significantly decreased after the initial 5 min and nearly disappeared to form the wrinkled surface after 20 min. The initial granule formation proceeded on the surface between 20 and 40 min. Finally, smaller granules developed on the surface after 120 min, compared with that for 5 cm3·min–1. These observations illustrated the development process of granules on the surface. At a CO2 flow rate of 100 cm3·min–1, as shown in Figure 6, the wrinkled surface was formed during the initial 5 min. This led to restricted formation of K2CO3·1.5H2O and K4H2(CO3)3·1.5H2O. The initial granule growth on the surface, which was accompanied by disappearance of the pores at 20 min (Figure 6f), caused the slow transformation process of K2CO3·1.5H2O and

ACS Paragon Plus Environment

12

Page 13 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

K4H2(CO3)3·1.5H2O between 5 and 20 min (Figure 4b and c). During this transformation period, the surface resistance continued to be enhanced due to the wrinkled surface formation covering the initial product particles. As a result, we observed that the K2CO3·1.5H2O phase remained at 100 cm3·min–1 and the formation of KHCO3 from K4H2(CO3)3·1.5H2O was limited already at 40 min (Figure 4d). Consequently, there was a faster decline in temperature, DTA, and DTG curves between 20 and 40 min, as seen in Figure 3, indicating that the CO2 occlusion rate at this CO2 flow rate was slower than that for other CO2 flow rates (with the exception of 1 cm3·min–1). The wrinkled surface must be broken in order to develop larger granules along with bicarbonate formation of K4H2(CO3)3·1.5H2O between 40 and 60 min so that the CO2 sorption rate and temperature can decrease according to the DTG and DTA curves in Figure 3. Considering the morphology and XRD patterns at 120 min for a CO2 flow rate of 100 cm3·min–1, we can see that the smaller granules on the external surface developed with the bicarbonate formation of K4H2(CO3)3·1.5H2O from 60 to 120 min, as seen in Figure 4e and Figure 2. Since smaller granules promote the chemisorption and gas diffusion processes,28, 29 the acceleration of reaction (8) between 60 and 80 min (final stage) was found to cause the observed sharp peaks in the DTA and DTG curves in Figure 3. To determine why the second peak appears for high CO2 flow rates (20–100 cm3·min–1) during the final stage in Figure 3, we compared the morphology at 120 min for a CO2 flow rate of 100 cm3·min–1 with that for a CO2 flow rate of 30 cm3·min–1. From this, it was found that smaller granules of KHCO3 developed on the surface at 120 min for the CO2 flow rates of 30 and 100 cm3·min–1 with the bicarbonate formation of K4H2(CO3)3·1.5H2O. Furthermore, the CO2 occlusion rate and accompanying temperature showed a transient and sharp increase during the final stage. In contrast, the morphology in the initial stage of CO2 occlusion played a curial role

ACS Paragon Plus Environment

13

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 24

in the overall reaction rate, which can account for the decline in temperature variation and the intensity of the peaks in the DTA and DTG curves, shown in Figure 3d–f, upon increasing CO2 flow rate with the exception of 1 cm3·min–1. 4. CONCLUSIONS The analysis of variations in XRD patterns with reaction time of the bicarbonate formation process of K2CO3 under various CO2 concentrations, illustrates that the reaction of K2CO3 with CO2 and H2O proceeds via three reactions. The results of changes in exothermic properties, temperature, and CO2 occlusion rate show that the CO2 occlusion process of K2CO3 also goes through three stages. The first product is K2CO3·1.5H2O, which can be converted to K4H2(CO3)3·1.5H2O in the initial stage of CO2 occlusion. The formation of KHCO3 from K4H2(CO3)3·1.5H2O proceeds slowly during the final stage, though there is a second increase in CO2 occlusion rate and temperature under high CO2 flow rates (20–100 cm3·min–1). The analysis of morphology variation with reaction time demonstrates that the CO2 occlusion of K2CO3 is inhibited at high CO2 concentrations and the formation of KHCO3 from K4H2(CO3)3·1.5H2O is favorable for high amounts of CO2 occlusion with increasing CO2 concentration. Particle morphology is found to play a crucial role in the kinetic behavior of CO2 sorptivity of K2CO3. TG–DTA and XRD analysis show the dependence of CO2 sorptivity of K2CO3 on the CO2 concentration under various CO2 atmospheres in the presence of saturated water vapor. Thus, K2CO3 from the decomposition of KHCO3 can be used as a sorbent to capture CO2 at relatively low temperatures. AUTHOR INFORMATION Corresponding Author

ACS Paragon Plus Environment

14

Page 15 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

*E-mail: [email protected]. Tel: +81(0)43–290–2784. Author Contributions The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript. Notes The authors declare no competing financial interest. ACKNOWLEDGMENT This work was supported by The Iwatani Naoji Foundation's Research Grant and partially by Grants-in-Aid for Challenging Exploratory Research and for Fundamental Scientific Research from the Japan Society for the Promotion of Science. The authors would like to thank Enago (www.enago.jp) for the English language review. REFERENCES (1) Aresta, M. Carbon dioxide recovery and utilization, Kluwer Academic Pub., Boston, 2003, 53. (2) Metz, B.; Davidson, O.; de Coninck, H.; Loos, M.; Meyer, L. IPCC Special Report on Carbon Dioxide Capture and Storage, Cambridge University Press, New York, 2005. (3) Wall, T. F. Proc. Combust. Inst. 2007, 31, 31–47. (4) Yang, W. C.; Hoffman, J. Ind. Eng. Chem. Res. 2009, 48, 341–351. (5) Fricker, K. J.; Park, A.–H. A. Eng.Sci. 2013, 100, 332–341.

ACS Paragon Plus Environment

15

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 24

(6) Plaza, M. G.; García, S.; Rubiera, F.; Pis, J. J.; Pevida, C. J.Chem. Eng. 2010, 163, 41–47. (7) Cavenati, S.; Grande, C. A.; Rodrigues, A. E. J. Chem. Eng. Data 2004, 49, 1095–1101. (8) Presser, V.; McDonough, J.; Yeon, S.H.; Gogotsi, Y. Energy Environ. Sci. 2011, 4, 3059– 3066. (9) Zhao, C.; Chen, X.; Zhao, C. Chemosphere 2009, 75, 1401–1404. (10) Park, Y. C.; Jo, S. H.; Ryu, C. K.; Yi; C. K. Energy Proc. 2011, 4, 1508–1512. (11) Park, Y. C.; Jo, S. H.; Lee, D. H.; Yi, C. K.; Ryu, C. K.; Kim, K. S.; You, C. H.; Park, Y. S. Energy Proc. 2013, 37, 122–126. (12) Park, S. W.; Sung, D. H.; Choi, B. S.; Lee, J. W.; Kumazawa, H. J. Ind. Eng. Chem. 2006, 12, 522–530. (13) Seo, Y.; Jo, S. H.; Ryu, C. K.; Yi, C. K. J. Environ. Eng. 2009, 135, 473–477. (14) Zhao, C.; Chen, X.; Zhao, C. Energ. Fuel 2009, 23, 4683–4687. (15) Zhao, C.; Chen, X.; Zhao, C. Int. J. Greenhouse Gas Control 2010, 4, 655–658. (16) Zhao, C.; Chen, X.; Zhao, C. Ind. Eng. Chem. Res. 2010, 49, 12212–12216. (17) Zhao, C.; Chen, X.; Zhao, C. Energ. Fuel 2012, 26, 1401–1405. (18) Hayashi, H.; Taniuchi, J.; Furuyashiki, N.; Sugiyama, S.; Hirano, S.; Shigemoto, N.; Nonaka, T. Ind. Eng. Chem. Res. 1998, 37, 185–191.

ACS Paragon Plus Environment

16

Page 17 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

(19) Shigemoto, N.; Yanagihara, T.; Sugiyama, S.; Hayashi, H. Energ. Fuels 2006, 20, 721– 726. (20) Seo, Y. W.; Jo, S. H.; Ryu, H. J.; Bae, D. H.; Ryu, C. K.; Yi, C. K. Korean J. Chem. Eng. 2007, 24, 457–460. (21) Lee, S. C.; Chae, H. J.; Choi, B. Y.; Jung, S. Y.; Ryu, C. Y.; Park, J. J.; Baek, J. I.; Ryu, C. K.; Kim, J. C. Korean J. Chem. Eng. 2011, 28, 480–486. (22) Chioyama, H.; Luo, H.; Ohba, T.; Kanoh, H. Adsorption Sci. and Technol. 2015, 33, 243– 250. (23) Zhao, C.; Guo, Y.; Li, C.; et al. Chem. Eng. J., 2014, 254, 524–530. (24) Zhao, C.; Chen, X.; Zhao, C.; Liu, Y. Energ. Fuel 2009, 23, 1766–1769. (25) Zhao, C.; Chen, X.; Zhao, C. Ind. Eng. Chem. Res. 2012, 51, 14361–14366. (26) Lee, S. C.; Choi, B. Y.; Ryu, C. K.; Ahn, Y. S.; Lee, T. J.; Kim, J. C. Korean J. Chem. Eng. 2006, 23, 374–379. (27) Lee, S. C.; Kim, J. C. Catal. Surv. Asia. 2007, 11, 171–185. (28) Venegas, M. J.; Fregoso–Israel, E.; Escamilla, R.; Pfeiffer, H. Ind. Eng. Chem. Res. 2007, 46, 2407–2412. (29) García–Labiano, F.; Abad, A.; de Diego, L .F.; Gayán, P.; Adánez, J. Chem. Eng. Technol. 2002, 57, 2381–2393.

ACS Paragon Plus Environment

17

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 24

Figure Captions Figure 1. TG curves (a) for bicarbonate formation under various CO2 flow rates (cm3·min–1), which are indicated with numbers, with N2 and H2O and magnified for 0–30 min (b). Figure 2. XRD patterns after CO2 occlusion of K2CO3 at various CO2 flow rates (cm3·min–1), which are indicated with numbers. ▼: KHCO3; ◇: K4H2(CO3)3·1.5H2O.

Figure 3. Temperature change (a), DTA (b), and DTG (c) curves at various CO2 flow rates (cm3·min–1), which are indicated with numbers. These figures are magnified for reaction times of 0–30 min in (d), (e), and (f), respectively. Figure 4. XRD patterns after CO2 occlusion of K2CO3 at various CO2 flow rates (cm3·min–1), which are indicated with numbers, with reaction time: (a) 1 min, (b) 5 min, (c) 20 min, (d) 40 min, and (e) 60 min. ▼: KHCO3; ◇: K4H2(CO3)3·1.5H2O; ◆: K2CO3·1.5H2O.

Figure 5. SEM images of K2CO3 before and after CO2 occlusion at a CO2 flow rate of 5 cm3·min–1 (a–d) (a: 0 min, b: 5 min, c: 20 min, and d: 120 min). Figure 6. SEM images of K2CO3 after CO2 occlusion at a CO2 flow rate of 100 cm3·min–1 (e–g) (e: 5 min, f: 20 min, and g: 120 min).

ACS Paragon Plus Environment

18

Page 19 of 24

100

a

10

90

100

20 30

5 10 20 30

b

5

95

85 1

90

100 1

Weight / %

Weight / %

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

80

85 80

75 75 70

70 0

20

40

60

80

100

120

0

5

Time / min

10

15

20

25

30

Time / min

Figure 1. TG curves (a) for bicarbonate formation under various CO2 flow rates (cm3·min–1), which are indicated with numbers, with N2 and H2O, and magnified for 0–30 min (b).

ACS Paragon Plus Environment

19

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

In ten sity / a.u

Energy & Fuels

▼▼ ▼ ◇▼



▼ ▼ ▼ 100 ▼

▼▼ ▼ ◇▼



▼ ▼ ▼

30 ▼

▼▼ ▼ ◇▼



▼ ◇ ▼◇ ▼

20 ▼

▼ ◇▼◇



◇ ▼ ◇▼ ▼ 10 ▼

▼◇▼◇



◇ ▼ ◇▼ ▼

5



▼ ▼◇ ◇ ◇



◇ ▼ ◇▼ ▼

1



◇ ▼ ◇

25

Page 20 of 24

30

35

40

45

2θ / degree

Figure 2. XRD patterns after CO2 occlusion of K2CO3 at various CO2 flow rates (cm3·min–1), which are indicated with numbers. ▼: KHCO3; ◇: K4H2(CO3)3·1.5H2O.

ACS Paragon Plus Environment

20

Page 21 of 24

318

10 20

5

5

d

10

e

317

316

Temperature / K

Temperature / K

318

a

317

10 20

30

315

100 1

314

30

316 100

315 314

313

313

312 25

312 25

b

1

20 30 5

20

15

DTA / a. u

DTA / a. u

20 30

10 20

100

10

1

15

100

10

1

5

5

5

0.8

d(weight) / dt / % min

-1

c

0 1.0

5

0.6

10 20 30

1

100

0.4

10

-1

0 1.0

d(weight) / dt / % min

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

0.2

0.8

f

5

20

30

100

0.6 1

0.4 0.2 0.0

0.0 0

20

40

60

80

100

120

0

5

Time / min

10

15

20

25

30

Time / min

Figure 3. Temperature change (a), DTA (b), and DTG (c) curves at various CO2 flow rates (cm3·min–1), which are indicated with numbers. These figures are magnified for reaction times of 0–30 min in (d), (e), and (f), respectively.

ACS Paragon Plus Environment

21

Energy & Fuels



◆ ◇◆ ◇◇ ◇

◆ ◆◆◆◆



◆ ◇◇ ◆ ◇

◆◆◆◆◆





◆◆ ◇◇ ◇



◆◆◆◆◆◆



◆ ◆ ◇◇ ◇



◆ ◆◆◆◆ ◆

25

30

b

Intensity / a.u



◆ ◇

◆ ◆ ◇ ◇◇





◆ ◇



◇ ▼ ◆

25

◇ ◆◇

30 ◆◆ ◆ ◇ ◆◆ ◆ ◇

◇ ◆◇

◆ 5 ◆ ◆◆◆ ◆



100 ◇

◇ ◇



◇ ◇▼

30

▼ ▼ ◆◆ ◇◇ ◆ ◇◇◇ ◇ ◇



◇◇

5



◇ ◇ ◆

1

◇◇ ▼▼ ◆ ◇ ◇◇ ◇◇









e ◇▼

100 ▼ ◇ ◇◇ ▼ ▼ ◇

▼ ▼

▼ ▼ ◇ ◇

▼ ◇

◇▼

◇◇

30

◇ ◇▼

◇ ◇

◇ ▼ ▼◆ ◆ ◇◇◇ ◆◇





◇◇



▼ ◆ ◇ ◇◇ ▼

◆◆◆◆◆◆ 1 ◇



◇◇

c

100 ◇◆◆◆ ◇ ◆ ◇ ◇◆ ◆ ◇

◇◇

◇◇ ▼▼ ▼◇



45

◆ ◆



◆◆

1

40

◆ ◆ ◇ ◇◇





35

5

◆ ◇◇◆

◆ ◇

d

30

Intensity / a.u



100

◆ ◆ ◇◇

▼ ◇ ▼◇

▼ 30

▼ ▼ ◇ ▼ ◇◇ ◇



5



40

45





100



▼ ▼▼▼ ◇ ▼◇ ▼



◇ ◇



◇ 1 ◇ ◆ ◆ ◇ ◆◆

35

Intensity / a.u

Intnesity / a.u

a

Intensity / a.u

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 24

25





30

2θ / degree

◇ ◆◆ ◇◇

35

2θ / degree

◇ ◇

40

1

◆◆ ◇

45

Figure 4. XRD patterns after CO2 occlusion of K2CO3 at various CO2 flow rates (cm3·min–1), which are indicated with numbers, with reaction time: (a) 1 min, (b) 5 min, (c) 20 min, (d) 40 min, and (e) 60 min. ▼: KHCO3; ◇: K4H2(CO3)3·1.5H2O; ◆: K2CO3·1.5H2O.

ACS Paragon Plus Environment

22

Page 23 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

Figure 5. SEM images of K2CO3 before and after CO2 occlusion at a CO2 flow rate of 5 cm3·min–1 (a–d) (a: 0 min, b: 5 min, c: 20 min, and d: 120 min).

ACS Paragon Plus Environment

23

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 24

Figure 6. SEM images of K2CO3 after CO2 occlusion at a CO2 flow rate of 100 cm3·min–1 (e–g) (e: 5 min, f: 20 min, and g: 120 min).

ACS Paragon Plus Environment

24